Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\setcounter{chapter}{-1} | |
\chapter{Categories} | |
\label{categorychapter} | |
The language of categories is not strictly necessary to understand the basics | |
of commutative | |
algebra. Nonetheless, it is extremely convenient and powerful. It will clarify | |
many of the constructions made in the future when we can freely use terms like | |
``universal property'' or ``adjoint functor.'' As a result, we begin this book | |
with a brief introduction to category theory. We only scratch the surface; the | |
interested reader can pursue further study in \cite{Ma98} or \cite{KS05}. | |
Nonetheless, the reader is advised not to take the present chapter too | |
seriously; skipping it for the moment to chapter 1 and returning here as a | |
reference could be quite reasonable. | |
\section{Introduction} | |
\subsection{Definitions} | |
Categories are supposed to be places where mathematical objects live. | |
Intuitively, to any given type of structure (e.g. groups, rings, etc.), | |
there should be a | |
category of objects with that structure. These are not, of course, the only | |
type of categories, but they will be the primary ones of concern to us in this | |
book. | |
The basic idea of a category is that there should be objects, and that one | |
should be able to map between objects. These mappings could be functions, and | |
they often are, but they don't have to be. Next, one has to be able to compose | |
mappings, and associativity and unit conditions are required. Nothing else is required. | |
\begin{definition} | |
A \textbf{category} $\mathcal{C}$ consists of: | |
\begin{enumerate} | |
\item A collection of \textbf{objects}, | |
$\ob \mathcal{C}$. | |
\item For each pair of objects $X, Y \in | |
\ob \mathcal{C}$, a set | |
of \textbf{morphisms} $\hom_{\mathcal{C}}(X, Y)$ (abbreviated $\hom(X,Y)$). | |
\item For each object $X \in \ob\mathcal{C}$, there is an \textbf{identity | |
morphism} | |
$1_X \in \hom_{\mathcal{C}}(X, X)$ (often just abbreviated to $1$). | |
\item There is a \textbf{composition law} | |
$\circ: \hom_{\mathcal{C}}(X, Y) \times \hom_{\mathcal{C}}(Y, Z) \to | |
\hom_{\mathcal{C}}(X, Z), (g, f) \to g | |
\circ f$ for every | |
triple $X, Y, Z$ of objects. | |
\item The composition law is unital and associative. | |
In other words, if $f \in \hom_{\mathcal{C}}(X, Y)$, then $1_Y \circ f = f | |
\circ 1_X = f$. Moreover, if $g \in \hom_{\mathcal{C}}(Y, Z)$ and $h \in | |
\hom_{\mathcal{\mathcal{C}}}(Z, W)$ for objects $Z, Y, W$, then | |
\[ h \circ (g \circ f) = (h \circ g) \circ f \in \hom_{\mathcal{C}}(X, W). \] | |
\end{enumerate} | |
\end{definition} | |
We shall write $f: X \to Y$ to denote an element of $\hom_{\mathcal{C}}(X, Y)$. | |
In practice, $\mathcal{C}$ will often be the storehouse for mathematical objects: groups, Lie algebras, | |
rings, etc., in which case these ``morphisms'' will just be ordinary functions. | |
Here is a simple list of examples. | |
\begin{example}[Categories of structured sets] | |
\begin{enumerate} | |
\item $\mathcal{C} = \mathbf{Sets}$; the objects are sets, and the morphisms | |
are functions. | |
\item $\mathcal{C} = \mathbf{Grps}$; the objects are groups, and the morphisms | |
are maps of groups (i.e. homomorphisms). | |
\item $\mathcal{C} = \mathbf{LieAlg}$; the objects are Lie algebras, and the | |
morphisms are maps of Lie algebras (i.e. homomorphisms).\footnote{Feel free to | |
omit if the notion of Lie algebra is unfamiliar.} | |
\item $\mathcal{C} = \mathbf{Vect}_k$; the objects are vector spaces over a | |
field $k$, and the morphisms are linear maps. | |
\item $\mathcal{C} = \mathbf{Top}$; the objects are topological spaces, and the | |
morphisms are continuous maps. | |
\item This example is slightly more subtle. Here the category $\mathcal{C}$ | |
has objects consisting of topological spaces, but the morphisms between two | |
topological spaces $X,Y$ are the \emph{homotopy classes} of maps $X \to Y$. | |
Since composition respects homotopy classes, this is well-defined. | |
\end{enumerate} | |
\end{example} | |
In general, the objects of a category do not have to form a set; they can | |
be too large for | |
that. | |
For instance, the collection of objects in $\mathbf{Sets}$ does not form a set. | |
\begin{definition} | |
A category is \textbf{small} if the collection of objects is a set. | |
\end{definition} | |
The standard examples of categories are the ones above: structured sets | |
together with structure-preserving maps. Nonetheless, one can easily give | |
other examples that are not of this form. | |
\begin{example}[Groups as categories] \label{BG} | |
Let $G$ be a finite group. Then we can make a category $B_G$ where the objects | |
just consist of one element $\ast$ and the maps $\ast \to \ast$ are the elements | |
$g \in G$. The identity is the identity of $G$ and composition is multiplication | |
in the group. | |
In this case, the category does not represent much of a class of objects, but | |
instead we think of the composition law as the key thing. So a group is a | |
special kind of (small) category. | |
\end{example} | |
\begin{example}[Monoids as categories] | |
A monoid is precisely a category with one object. Recall that a \textbf{monoid} | |
is a set together with an associative and unital multiplication (but which | |
need not have inverses). | |
\end{example} | |
\begin{example}[Posets as categories] \label{posetcategory} Let $(P, \leq)$ be a partially ordered | |
(or even preordered) set (i.e. poset). Then $P$ can be regarded as a (small) category, where the objects are the elements | |
$p \in P$, and $$\hom_P(p, q) = \begin{cases} | |
\ast & \text{if } p \leq q \\ | |
\emptyset & \text{otherwise} | |
\end{cases} $$ | |
\end{example} | |
There is, however, a major difference between category theory and set theory. | |
There is \textbf{nothing} in the language of categories that lets one look | |
\emph{inside} an object. We think of vector spaces having elements, spaces | |
having points, etc. | |
By contrast, categories treat these kinds of things as invisible. There | |
is nothing ``inside'' of an object $X \in \mathcal{C}$; the only way to | |
understand $X$ is | |
to understand the ways one can map into and out of $X$. | |
Even if one is working with a category of ``structured sets,'' the underlying | |
set of an object in this category is not part of the categorical data. | |
However, there are instances in which the ``underlying set'' can be recovered | |
as a $\hom$-set. | |
\begin{example} | |
In the category $\mathbf{Top}$ of topological spaces, one can in fact recover the | |
``underlying set'' of a topological space via the hom-sets. Namely, for each | |
topological space, the points of $X$ are the same thing as the mappings from a | |
one-point space into $X$. | |
That is, we have | |
\[ |X| = \hom_{\mathbf{Top}}(\ast, X), \] | |
where $\ast$ is the one-point space. | |
Later we will say that the functor assigning to each | |
space its underlying set is \emph{corepresentable.} | |
\end{example} | |
\begin{example} | |
Let $\mathbf{Ab}$ be the category of abelian groups and group-homomorphisms. Again, the claim is that | |
using only this category, one can recover the underlying set of a given abelian | |
group $A$. This is because the elements of $A$ can be canonically identified | |
with \emph{morphisms} $\mathbb{Z} \to A$ (based on where $1 \in \mathbb{Z}$ | |
maps). | |
\end{example} | |
\begin{definition} | |
We say that $\mathcal{C}$ is a \textbf{subcategory} of the category | |
$\mathcal{D}$ if the collection of objects of $\mathcal{C}$ is a subclass of | |
the collection of objects of $\mathcal{D}$, and if whenever | |
$X, Y \in \mathcal{C}$, we have | |
\[ \hom_{\mathcal{C}}(X, Y) \subset \hom_{\mathcal{D}}(X, Y) \] | |
with the laws of composition in $\mathcal{C}$ induced by that in $\mathcal{D}$. | |
$\mathcal{C}$ is called a \textbf{full subcategory} if $\hom_{\mathcal{C}}(X, | |
Y) = \hom_{\mathcal{D}}(X, Y)$ whenever $X, Y \in \mathcal{C}$. | |
\end{definition} | |
\begin{example} | |
The category of abelian groups is a full subcategory of the category of groups. | |
\end{example} | |
\subsection{The language of commutative diagrams} | |
While the language of categories is, of course, purely algebraic, it will be | |
convenient for psychological reasons to visualize categorical arguments | |
through diagrams. | |
We shall introduce this notation here. | |
Let $\mathcal{C}$ be a category, and let $X, Y$ be objects in $\mathcal{C}$. | |
If $f \in \hom(X, Y)$, we shall sometimes write $f$ as an arrow | |
\[ f: X \to Y \] | |
or | |
\[ X \stackrel{f}{\to} Y \] | |
as if $f$ were an actual function. | |
If $X \stackrel{f}{\to} Y$ and $Y \stackrel{g}{\to} Z$ are morphisms, | |
composition $g \circ f: X \to Z$ can be visualized by the picture | |
\[ X \stackrel{f}{\to} Y \stackrel{g}{\to} Z.\] | |
Finally, when we work with several objects, we shall often draw collections of | |
morphisms into diagrams, where arrows indicate morphisms between two objects. | |
\begin{definition} | |
A diagram will be said to \textbf{commute} if whenever one goes from one | |
object in the diagram to another by following the arrows in the right order, | |
one obtains the same morphism. | |
For instance, the commutativity of the diagram | |
\[ \xymatrix{ | |
X \ar[d]^f \ar[r]^{f'} & W \ar[d]^g \\ | |
Y \ar[r]^{g'} & Z | |
}\] | |
is equivalent to the assertion that | |
\[ g \circ f' = g' \circ f \in \hom(X, Z). \] | |
\end{definition} | |
As an example, the assertion that the associative law holds in a category | |
$\mathcal{C}$ can be stated as follows. For every quadruple $X, Y, Z, W \in | |
\mathcal{C}$, the following diagram (of \emph{sets}) commutes: | |
\[ \xymatrix{ | |
\hom(X, Y) \times \hom(Y, Z) \times \hom(Z, W) \ar[r] \ar[d] & \hom(X, Z) | |
\times \hom(Z, W) \ar[d] \\ | |
\hom(X,Y) \times \hom(Y, W) \ar[r] & \hom(X, W). | |
}\] | |
Here the maps are all given by the composition laws in $\mathcal{C}$. | |
For instance, the downward map to the left is the product of the identity on | |
$\hom(X, Y)$ with the composition law $\hom(Y, Z) \times \hom(Z, W) \to \hom(Y, | |
W)$. | |
\subsection{Isomorphisms} | |
Classically, one can define an isomorphism of groups as a bijection that | |
preserves the group structure. This does not generalize well to categories, as | |
we do not have a notion of ``bijection,'' as there is no way (in general) to | |
talk about the ``underlying set'' of an object. | |
Moreover, this definition does not generalize well to topological spaces: | |
there, an isomorphism should not just be a bijection, but something which | |
preserves the topology (in a strong sense), i.e. a homeomorphism. | |
Thus we make: | |
\begin{definition} | |
An \textbf{isomorphism} between objects $X, Y$ in a category $\mathcal{C}$ is a | |
map $f: X \to Y$ such that there exists $g: Y \to X$ with | |
\[ g \circ f = 1_X, \quad f \circ g = 1_Y. \] | |
Such a $g$ is called an \textbf{inverse} to $f$. \end{definition} | |
\begin{remark} | |
It is easy to check that the inverse $g$ is | |
unique. Indeed, suppose $g, g'$ both were inverses to $f$. Then | |
\[ g' = g' \circ 1_Y = g' \circ (f \circ g) = (g' \circ f) \circ g = 1_X | |
\circ g = g. \] | |
\end{remark} | |
This notion is isomorphism is more correct than the idea of being one-to-one and onto. A bijection of | |
topological spaces is not necessarily a homeomorphism. | |
\begin{example} | |
It is easy to check that an isomorphism in the category $\mathbf{Grp}$ is an | |
isomorphism of groups, that an isomorphism in the category $\mathbf{Set}$ is a | |
bijection, and so on. | |
\end{example} | |
We are supposed to be able to identify isomorphic objects. In the categorical | |
sense, this means mapping into $X$ should be the same as mapping into $Y$, if | |
$X, Y$ are isomorphic, via an isomorphism $f: X \to Y$. | |
Indeed, let | |
$Z$ be another object of $\mathcal{C}$. | |
Then we can define a map | |
\[ \hom_{\mathcal{C}}(Z, X) \to \hom_{\mathcal{C}}(Z, Y) \] | |
given by post-composition with $f$. This is a \emph{bijection} if $f$ is an | |
isomorphism (the inverse is given by postcomposition with the inverse to $f$). | |
Similarly, one can easily see that mapping \emph{out of} $X$ is essentially the | |
same as mapping out of $Y$. | |
Anything in general category theory that is true for $X$ should be true for $Y$ | |
(as general category theory can only try to understand $X$ in terms of maps | |
into or out of it!). | |
\begin{exercise} | |
The relation ``$X, Y$ are isomorphic'' is an equivalence relation on the class | |
of objects of a category $\mathcal{C}$. | |
\end{exercise} | |
\begin{exercise} | |
Let $P$ be a preordered set, and make $P$ into a category as in | |
\cref{posetcategory}. Then $P$ is a poset if and only if two isomorphic objects | |
are equal. | |
\end{exercise} | |
For the next exercise, we need: | |
\begin{definition} | |
A \textbf{groupoid} is a category where every morphism is an isomorphism. | |
\end{definition} | |
\begin{exercise} | |
The | |
sets $\hom_{\mathcal{C}}(A, A)$ are \emph{groups} if $\mathcal{C}$ is a | |
groupoid and $A \in \mathcal{C}$. A group is essentially the same as a groupoid | |
with one object. | |
\end{exercise} | |
\begin{exercise} | |
Show that the following is a groupoid. Let $X$ be a topological space, and let | |
$\Pi_1(X)$ be the category defined as follows: the objects are elements of $X$, | |
and morphisms $x \to y$ (for $x,y \in X$) are homotopy classes of maps $[0,1] | |
\to X$ (i.e. paths) that send $0 \mapsto x, 1 \mapsto y$. Composition of maps | |
is given by concatenation of paths. | |
(Check that, because one is working with \emph{homotopy classes} of paths, | |
composition is associative.) | |
$\Pi_1(X)$ is called the \textbf{fundamental groupoid} of $X$. Note that | |
$\hom_{\Pi_1(X)}(x, x)$ is the \textbf{fundamental group} $\pi_1(X, x)$. | |
\end{exercise} | |
\section{Functors} | |
A functor is a way of mapping from one category to another: each object is sent | |
to another object, and each morphism is sent to another morphism. We shall | |
study many functors in the sequel: localization, the tensor product, $\hom$, | |
and fancier ones like $\tor, \ext$, and local cohomology functors. | |
The main benefit of a functor is that it doesn't simply send objects to | |
other objects, but also morphisms to morphisms: this allows one to get new commutative | |
diagrams from old ones. | |
This will turn out to be a powerful tool. | |
\subsection{Covariant functors} | |
Let $\mathcal{C}, \mathcal{D}$ be categories. If $\mathcal{C}, \mathcal{D}$ | |
are categories of structured sets (of possibly different types), there may be a | |
way to associate objects in $\mathcal{D}$ to objects in $\mathcal{C}$. For | |
instance, to every group $G$ we can associate its \emph{group ring} | |
$\mathbb{Z}[G]$ | |
(which we do not define here); to each topological space we can associate its | |
\emph{singular chain complex}, and so on. | |
In many cases, given a map between objects in $\mathcal{C}$ preserving the | |
relevant structure, there will be an induced map on the corresponding objects | |
in $\mathcal{D}$. It is from here that we define a \emph{functor.} | |
\begin{definition} \label{covfunc} | |
A \textbf{functor} $F: \mathcal{C} \to \mathcal{D}$ consists of a function $F: | |
\mathcal{C} \to \mathcal{D}$ (that is, a rule that assigns to each object | |
in $\mathcal{C}$ an object of $\mathcal{D}$) and, for each pair $X, Y \in | |
\mathcal{C}$, | |
a map | |
$F: \hom_{\mathcal{C}}(X, Y) \to \hom_{\mathcal{D}}(FX, FY)$, which preserves | |
the identity | |
maps and composition. | |
In detail, the last two conditions state the following. | |
\begin{enumerate} | |
\item If $X \in | |
\mathcal{C}$, then $F(1_X)$ is the identity morphism $1_{F(X)}: F(X) \to | |
F(X)$. | |
\item If $A \stackrel{f}{\to} B \stackrel{g}{\to} C$ are | |
morphisms in $\mathcal{C}$, | |
then $F(g \circ f) = F(g) \circ F(f)$ as morphisms $F(A) \to F(C)$. | |
Alternatively, we can say that $F$ \emph{preserves commutative diagrams.} | |
\end{enumerate} | |
\end{definition} | |
In the last statement of the definition, note that if | |
\[ \xymatrix{ | |
X \ar[rd]^h \ar[r]^f & Y \ar[d]^g \\ | |
& Z | |
}\] | |
is a commutative diagram in $\mathcal{C}$, then the diagram obtained by | |
applying the functor $F$, namely | |
\[ \xymatrix{ | |
F(X) \ar[rd]^{F(h)} \ar[r]^{F(f)} & F(Y) \ar[d]^{F(g)} \\ | |
& F(Z) | |
}\] | |
also commutes. It follows that applying $F$ to more complicated commutative | |
diagrams also yields new commutative diagrams. | |
Let us give a few examples of functors. | |
\begin{example} | |
There is a functor from $\mathbf{Sets} \to \mathbf{AbelianGrp}$ sending a set | |
$S$ to the free abelian group on the set. (For the definition of a free abelian | |
group, or more generally a free $R$-module over a ring $R$, see | |
\cref{freemoduledef}.) | |
\end{example} | |
\begin{example} \label{pi0} | |
Let $X$ be a topological space. Then to it we can associate the set $\pi_0(X)$ | |
of \emph{connected components} of $X$. | |
Recall that the continuous image of a | |
connected set is connected, so if $f: X \to Y$ is a continuous map and $X' | |
\subset X$ connected, $f(X')$ is contained in a connected component of $Y$. It | |
follows that $\pi_0$ is a functor $\mathbf{Top} \to \mathbf{Sets}$. | |
In fact, it is a functor on the \emph{homotopy category} as well, because | |
homotopic maps induce the same maps on $\pi_0$. | |
\end{example} | |
\begin{example} | |
Fix $n$. | |
There is a functor from $\mathbf{Top} \to \mathbf{AbGrp}$ | |
(categories of topological spaces and abelian groups) sending a | |
space $X$ to its $n$th homology group $H_n(X)$. We know that given a map of spaces | |
$f: X \to Y$, | |
we get a map of abelian groups $f_*: H_n(X) \to H_n(Y)$. See \cite{Ha02}, for | |
instance. | |
\end{example} | |
We shall often need to compose functors. For instance, we will want to see, for | |
instance, that the \emph{tensor product} (to be defined later, see | |
\cref{sec:tensorprod}) is | |
associative, which is really a statement about composing functors. The | |
following (mostly self-explanatory) definition elucidates this. | |
\begin{definition}\label{composefunctors} | |
If $\mathcal{C}, \mathcal{D}, \mathcal{E}$ are categories, $F: \mathcal{C} \to | |
\mathcal{D}, G: \mathcal{D} \to \mathcal{E}$ are covariant functors, then one | |
can define a \textbf{composite functor} | |
\[ F \circ G: \mathcal{C} \to \mathcal{E} \] | |
This sends an object $X \in \mathcal{C}$ to $G(F(X))$. | |
Similarly, a morphism $f :X \to Y$ is sent to $G(F(f)): G(F(X)) \to G(F(Y))$. | |
We leave the reader to check that this is well-defined. | |
\end{definition} | |
\begin{example}\label{categoryofcats} | |
In fact, because we can compose functors, there is a \emph{category of | |
categories.} Let $\mathbf{Cat}$ have objects as the small categories, and | |
morphisms as functors. Composition is defined as in \cref{composefunctors}. | |
\end{example} | |
\begin{example}[Group actions] \label{groupact} Fix a group $G$. | |
Let us understand what a functor $B_G \stackrel{}{\to} \mathbf{Sets}$ is. Here $B_G$ is the | |
category of \cref{BG}. | |
The unique object $\ast$ of $B_G$ goes to some set $X$. For each element $g \in G$, we | |
get a map $g: \ast \to \ast$ and thus a map $X \to X$. This is supposed to | |
preserve the composition law (which in $G$ is just multiplication), as well as | |
identities. | |
In particular, we get maps $i_g: X \to X$ corresponding to each $g \in G$, such | |
that the following diagram commutes for each $g_1, g_2 \in G$: | |
\[ \xymatrix{ | |
X \ar[r]^{i_{g_1}} \ar[rd]_{i_{g_1g_2}} & X \ar[d]^{i_{g_2}} \\ & X. | |
}\] | |
Moreover, if $e \in G$ is the identity, then $i_e = 1_X$. | |
So a functor $B_G \to \mathbf{Sets}$ is just a left $G$-action on a set $X$. | |
\end{example} | |
An important example of functors is given by the following. Let $\mathcal{C}$ | |
be a category of ``structured sets.'' Then, there is a functor $F: \mathcal{C} | |
\to \textbf{Sets}$ that sends a structured set to the underlying set. For | |
instance, there is a functor from groups to sets that forgets the group | |
structure. | |
More generally, suppose given two categories $\mathcal{C}, \mathcal{D}$, such | |
that $\mathcal{C}$ can be regarded as ``structured objects in $\mathcal{D}$.'' | |
Then there is a functor $\mathcal{C} \to \mathcal{D}$ that forgets the | |
structure. | |
Such examples are called \emph{forgetful functors.} | |
\subsection{Contravariant functors} | |
Sometimes what we have described above are called \textit{covariant functors}. | |
Indeed, we shall also be interested in similar objects that reverse the | |
arrows, such as duality functors: | |
\begin{definition} | |
A \textbf{contravariant functor} $\mathcal{C} | |
\stackrel{F}{\to}\mathcal{D}$ (between categories $\mathcal{C}, \mathcal{D}$) | |
is similar | |
data as in \cref{covfunc} except that now a map $X \stackrel{f}{\to} Y$ now | |
goes to a map $F(Y) | |
\stackrel{F(f)}{\to} F(X)$. Composites | |
are required to be preserved, albeit in the other direction. | |
In other words, if $X \stackrel{f}{\to} Y, Y \stackrel{g}{\to} Z$ are | |
morphisms, then we require | |
\[ F ( g \circ f) = F(f) \circ F(g): F(Z) \to F(X). \] | |
\end{definition} | |
We shall sometimes say just ``functor'' for \emph{covariant functor}. When we are | |
dealing with a contravariant functor, we will always say the word | |
``contravariant.'' | |
A contravariant functor also preserves commutative diagrams, except that the | |
arrows have to be reversed. For instance, if $F: \mathcal{C} \to \mathcal{D}$ | |
is contravariant and the diagram | |
\[ \xymatrix{ | |
A \ar[d] \ar[r] & C\\ | |
B \ar[ru] | |
}\] | |
is commutative in $\mathcal{C}$, then the diagram | |
\[ \xymatrix{ | |
F(A) & \ar[l] \ar[ld] F(C)\\ | |
F(B) \ar[u] | |
}\] | |
commutes in $\mathcal{D}$. | |
One can, of course, compose contravariant functors as in \cref{composefunctors}. But the composition of two | |
contravariant functors will be \emph{covariant.} So there is no ``category of | |
categories'' where the morphisms between categories are contravariant functors. | |
Similarly as in \cref{groupact}, we have: | |
\begin{example} | |
A \textbf{contravariant} functor from $B_G$ (defined as in \cref{BG}) to $\mathbf{Sets}$ corresponds to a | |
set with a \emph{right} $G$-action. | |
\end{example} | |
\begin{example}[Singular cohomology] | |
In algebraic topology, one encounters contravariant functors on the homotopy | |
category of topological spaces via the \emph{singular cohomology} functors $X | |
\mapsto H^n(X; \mathbb{Z})$. Given a continuous map $f: X \to Y$, there is a | |
homomorphism of groups | |
\[ f^* : H^n(Y; \mathbb{Z}) \to H^n(X; \mathbb{Z}). \] | |
\end{example} | |
\begin{example}[Duality for vector spaces] \label{dualspace} | |
On the category $\mathbf{Vect}$ of vector spaces over a field $k$, we | |
have | |
the contravariant functor | |
\[ V \mapsto V^{\vee}. \] | |
sending a vector space to its dual $V^{\vee} = \hom(V,k)$. | |
Given a map $V \to W$ of vector spaces, there is an induced map | |
\[ W^{\vee} \to V^{\vee} \] | |
given by the transpose. | |
\end{example} | |
\begin{example} | |
If we map $B_G \to B_G$ sending $\ast \mapsto \ast$ and $g \mapsto g^{-1}$, we | |
get a | |
contravariant functor. | |
\end{example} | |
We now give a useful (linguistic) device for translating between covariance and | |
contravariance. | |
\begin{definition}[The opposite category] \label{oppositecategory} | |
Let $\mathcal{C}$ be a category. Define the \textbf{opposite category} | |
$\mathcal{C}^{op}$ of $\mathcal{C}$ to have the same objects as | |
$\mathcal{C}$ but such that the morphisms between $X,Y$ in | |
$\mathcal{C}^{op}$ | |
are those between $Y$ and $X$ in $\mathcal{C}$. | |
\end{definition} | |
There is a contravariant functor $\mathcal{C} \to | |
\mathcal{C}^{op}$. | |
In fact, contravariant functors out of $\mathcal{C}$ are the \emph{same} as | |
covariant functors out of $\mathcal{C}^{op}$. | |
As a result, when results are often stated for both covariant and contravariant | |
functors, for instance, we can often reduce to the covariant case by using the | |
opposite category. | |
\begin{exercise} | |
A map that is an isomorphism in $\mathcal{C}$ corresponds to an isomorphism in | |
$\mathcal{C}^{op}$. | |
\end{exercise} | |
\subsection{Functors and isomorphisms} | |
Now we want to prove a simple and intuitive fact: if isomorphisms allow one to | |
say that one object in a category is ``essentially the same'' as another, | |
functors should be expected to preserve this. | |
\begin{proposition} | |
If $f: X \to Y$ is a map in $\mathcal{C}$, and $F: \mathcal{C} \to \mathcal{D}$ | |
is a functor, then $F(f): FX \to FY$ is an isomorphism. | |
\end{proposition} | |
The proof is quite straightforward, though there is an important point here. | |
Note that the analogous result holds for \emph{contravariant} functors too. | |
\begin{proof} | |
If we have maps $f: X \to Y$ and $g : Y \to X$ such that the composites both | |
ways are identities, then we can apply the functor $F$ to this, and we find | |
that since | |
\[ f \circ g = 1_Y, \quad g \circ f = 1_X, \] | |
it must hold that | |
\[ F(f) \circ F(g) = 1_{F(Y)}, \quad F(g) \circ F(f) = 1_{F(X)}. \] | |
We have used the fact that functors preserve composition and identities. This | |
implies that $F(f)$ is an isomorphism, with inverse $F(g)$. | |
\end{proof} | |
Categories have a way of making things so general that are trivial. Hence, | |
this material is called general abstract nonsense. | |
Moreover, there is another philosophical point about category theory to | |
be made here: often, it is the definitions, and not the proofs, that matter. | |
For instance, what matters here is not the theorem, but the \emph{definition of | |
an | |
isomorphism.} It is a categorical one, and much more general than the usual | |
notion via injectivity and surjectivity. | |
\begin{example} | |
As a simple example, $\left\{0,1\right\}$ and $[0,1]$ are not isomorphic in the | |
homotopy category of topological spaces (i.e. are not homotopy equivalent) | |
because $\pi_0([0,1]) = \ast$ while $\pi_0(\left\{0,1\right\}) $ has two | |
elements. | |
\end{example} | |
\begin{example} | |
More generally, the higher homotopy group functors $\pi_n$ (see \cite{Ha02}) can be used to show | |
that the $n$-sphere $S^n$ is not homotopy equivalent to a point. For then | |
$\pi_n(S^n, \ast)$ would be trivial, and it is not. | |
\end{example} | |
There is room, nevertheless, for something else. Instead of having | |
something that sends objects to other objects, one could have something that | |
sends an object to a map. | |
\subsection{Natural transformations} | |
Suppose $F, G: \mathcal{C} \to \mathcal{D}$ are functors. | |
\begin{definition} | |
A \textbf{natural transformation} $T: F \to G$ consists of the following data. | |
For each $X \in C$, there is a morphism $TX: FX \to GX$ satisfying the | |
following | |
condition. Whenever $f: X \to Y$ is a morphism, the following diagram must | |
commute: | |
\[ \xymatrix{ | |
FX \ar[d]^{TX }\ar[r]^{F(f)} & FY \ar[d]^{TY} \\ | |
GX \ar[r]^{G(f)} & GY | |
}.\] | |
If $TX$ is an isomorphism for each $X$, then we shall say that $T$ is a | |
\textbf{natural isomorphism.} | |
\end{definition} | |
It is similarly possible to define the notion of a natural transformation | |
between \emph{contravariant} functors. | |
When we say that things are ``natural'' in the future, we will mean that the | |
transformation between functors is natural in this sense. | |
We shall use this language to state theorems conveniently. | |
\begin{example}[The double dual] | |
Here is the canonical example of ``naturality.'' | |
Let $\mathcal{C}$ be the category of finite-dimensional vector spaces over a | |
given field $k$. Let us further restrict the category such that the only | |
morphisms are the \emph{isomorphisms} of vector spaces. | |
For each $V \in \mathcal{C}$, we know that there is an isomorphism | |
\[ V \simeq V^{\vee} = \hom_k(V, k), \] | |
because both have the same dimension. | |
Moreover, the maps $V \mapsto V, V \mapsto V^{\vee}$ are both covariant functors on | |
$\mathcal{C}$.\footnote{Note that the dual $\vee$ was defined as a | |
\emph{contravariant} functor in \cref{dualspace}.} The first is the identity functor; for the second, if $f: V \to | |
W$ is an isomorphism, then there is induced a transpose map $f^t: W^{\vee} \to V^{\vee}$ | |
(defined by sending a map $W \to k$ to the precomposition $V \stackrel{f}{\to} | |
W \to k$), which is an isomorphism; we can take its inverse. | |
So we have two functors from $\mathcal{C}$ to itself, the identity and the | |
dual, and we know that $V \simeq V^{\vee}$ for each $V$ (though we have not | |
chosen any particular set of isomorphisms). | |
However, the isomorphism $V \simeq | |
V^{\vee}$ \emph{cannot} be made natural. That is, there is no way of choosing | |
isomorphisms | |
\[ T_V: V \simeq V^{\vee} \] | |
such that, | |
whenever $f: V \to W$ is an isomorphism of vector spaces, the following diagram | |
commutes: | |
\[ \xymatrix{ | |
V \ar[r]^{f} \ar[d]^{T_V} & W \ar[d]^{T_W} \\ | |
V^{\vee} \ar[r]^{(f^t)^{-1}} & W^{\vee}. | |
}\] | |
Indeed, fix $d>1$, and choose $V = k^d$. | |
Identify $V^{\vee}$ with $k^d$, and so the map $T_V$ is a $d$-by-$d$ matrix $M$ | |
with coefficients in $k$. The requirement is that for each \emph{invertible} | |
$d$-by-$d$ matrix $N$, we have | |
\[ (N^t)^{-1}M = MN, \] | |
by considering the above diagram with $V = W = k^d$, and $f$ corresponding to | |
the matrix $N$. | |
This is impossible unless $M = 0$, by elementary linear algebra. | |
Nonetheless, it \emph{is} possible to choose a natural isomorphism | |
\[ V \simeq V^{\vee \vee}. \] | |
To do this, given $V$, recall that $V^{\vee \vee}$ is the collection of maps | |
$V^{\vee} \to k$. To give a map $V \to V^{\vee \vee}$ is thus the same as | |
giving linear functions $l_v, v \in V$ such that $l_v: V \to k$ is linear in | |
$v$. We can do this by letting $l_v$ be ``evaluation at $v$.'' | |
That is, $l_v$ sends a linear functional $\ell: V \to k$ to $\ell(v) \in k$. We | |
leave it to the reader to check (easily) that this defines a homomorphism $V | |
\to V^{\vee \vee}$, and that everything is natural. | |
\end{example} | |
\begin{exercise} | |
Suppose there are two functors $B_G \to | |
\mathbf{Sets}$, i.e. $G$-sets. What is a natural transformation between them? | |
\end{exercise} | |
Natural transformations can be \emph{composed}. Suppose given functors $F, G, | |
H: \mathcal{C} \to \mathcal{D}$ a natural | |
transformation $T: F \to G$ and a natural transformation $U: G \to H$. | |
Then, for each $X \in \mathcal{C}$, we have maps $TX: FX \to GX, UX: GX \to | |
HY$. We can compose $U$ with $T$ to get a natural transformation $U \circ T: F | |
\to H$. | |
In fact, we can thus define a \emph{category} of functors | |
$\mathrm{Fun}(\mathcal{C}, \mathcal{D})$ (at least if $\mathcal{C}, | |
\mathcal{D}$ are small). The objects of this category are the functors $F: | |
\mathcal{C} \to \mathcal{D}$. The morphisms are natural transformations between | |
functors. Composition of morphisms is as above. | |
\subsection{Equivalences of categories} | |
Often we want to say that two categories $\mathcal{C}, \mathcal{D}$ are ``essentially the same.'' One way | |
of formulating this precisely is to say that $\mathcal{C}, \mathcal{D}$ are | |
\emph{isomorphic} in the category of categories. Unwinding the definitions, | |
this means that there exist functors | |
\[ F: \mathcal{C} \to \mathcal{D}, \quad G: \mathcal{D} \to \mathcal{C} \] | |
such that $F \circ G = 1_{\mathcal{D}}, G \circ F = 1_{\mathcal{C}}$. | |
This notion, of \emph{isomorphism} of categories, is generally far too | |
restrictive. | |
For instance, we could consider the category of all finite-dimensional vector | |
spaces over a given field $k$, and we could consider the full subcategory | |
of vector spaces of the form $k^n$. Clearly both categories encode essentially | |
the same mathematics, in some sense, but they are not isomorphic: one has a | |
countable set of objects, while the other has an uncountable set of objects. | |
Thus, we need a more refined way of saying that two categories are | |
``essentially the same.'' | |
\begin{definition} | |
Two categories $\mathcal{C}, \mathcal{D}$ are called \textbf{equivalent} if | |
there are functors | |
\[ F: \mathcal{C} \to \mathcal{D}, \quad G: \mathcal{D} \to \mathcal{C} \] | |
and natural isomorphisms | |
\[ F G \simeq 1_{\mathcal{D}}, \quad GF \simeq 1_{\mathcal{C}}. \] | |
\end{definition} | |
For instance, the category of all vector spaces of the form $k^n$ is equivalent | |
to the category of all finite-dimensional vector spaces. | |
One functor is the inclusion from vector spaces of the form $k^n$; the other | |
functor maps a finite-dimensional vector space $V$ to $k^{\dim V}$. Defining | |
the second functor properly is, however, a little more subtle. | |
The next criterion will be useful. | |
\begin{definition} | |
A functor $F: \mathcal{C} \to \mathcal{D}$ is \textbf{fully faithful} if $F: | |
\hom_{\mathcal{C}}(X, Y) \to \hom_{\mathcal{D}}(FX, FY)$ is a bijection for each pair of objects $X, Y \in | |
\mathcal{C}$. | |
$F$ is called \textbf{essentially surjective} if every element of $\mathcal{D}$ | |
is isomorphic to an object in the image of $F$. | |
\end{definition} | |
So, for instance, the inclusion of a full subcategory is fully faithful (by | |
definition). The forgetful functor from groups to sets is not fully faithful, | |
because not all functions between groups are automatically homomorphisms. | |
\begin{proposition} | |
A functor $F: \mathcal{C} \to \mathcal{D}$ induces an equivalence of categories | |
if and only if it is fully faithful and essentially surjective. | |
\end{proposition} | |
\begin{proof} | |
\add{this proof, and the definitions in the statement.} | |
\end{proof} | |
\section{Various universal constructions} | |
Now that we have introduced the idea of a category and showed that a functor | |
takes isomorphisms to isomorphisms, we shall take various steps to characterize objects in terms of | |
maps (the most complete of which is the Yoneda lemma, \cref{yonedalemma}). In | |
general category | |
theory, this is generally all we \emph{can} do, since this is all the data we | |
are given. | |
We shall describe objects satisfying certain ``universal properties'' here. | |
As motivation, we first discuss the concept of the ``product'' in terms of a | |
universal property. | |
\subsection{Products} | |
Recall that if we have two sets $X$ and $Y$, the product $X\times Y$ is the set | |
of all elements of the form $(x,y)$ where $x\in X$ and $y\in Y$. The product is | |
also equipped with natural projections $p_1: X \times Y \to X$ and $p_2: X | |
\times Y \to Y$ that take $(x,y)$ to $x$ | |
and $y$ respectively. Thus any element of $X\times Y$ is uniquely determined by | |
where they project to on $X$ and $Y$. In fact, this is the case more generally; if | |
we have an index set $I$ and a product $X=\prod_{i\in I} X_i$, then an element | |
$x\in X$ determined uniquely by where where the projections $p_i(x)$ land in | |
$X_i$. | |
To get into the categorical spirit, we should speak not of elements but of maps | |
to $X$. Here is the general observation: if we have any other set $S$ with maps | |
$f_i:S\rightarrow X_i$ then there is a unique map $S\rightarrow X=\prod_{i\in | |
I}X_i$ given by sending $s\in S$ to the element $\{ f_i(s)\}_{i\in I}$. This | |
leads to the following characterization of a product using only ``mapping | |
properties.'' | |
\begin{definition} Let $\{X_i\}_{i\in I}$ be a collection of objects in some | |
category $\mathcal{C}$. Then an object $P \in \mathcal{C}$ with projections $p_i: P\rightarrow X_i$ | |
is said to be the \textbf{product} $\prod_{i\in I} X_i$ if the following ``universal | |
property'' holds: | |
let $S$ be any other object in $\mathcal{C}$ with maps $f_i:S\rightarrow X_i$. | |
Then there is a unique morphism $f:S\rightarrow P$ such that $p_i f = f_i$. | |
\end{definition} | |
In other words, to map into $X$ is the same as mapping into all the | |
$\left\{X_i\right\}$ at once. We have thus given a precise description of how | |
to map into $X$. | |
Note that, however, the product need not exist! | |
If it does, however, we can express the above formalism by the following | |
natural isomorphism | |
of contravariant functors | |
\[ \hom(\cdot, \prod_I X_i) \simeq \prod_I \hom(\cdot, X_i). \] | |
This is precisely the meaning of the last part of the definition. Note that | |
this observation shows that products in the category of \emph{sets} are really | |
fundamental to the idea of products in any category. | |
\begin{example} One of the benefits of this construction is that an actual | |
category is not specified; thus when we take $\mathcal{C}$ to be | |
$\mathbf{Sets}$, we | |
recover the cartesian product notion of sets, but if we take $\mathcal{C}$ to | |
be $\mathbf{Grp}$, we achieve the regular notion of the product of groups (the reader is | |
invited to check these statements). \end{example} | |
The categorical product is not unique, but it is as close to being so as | |
possible. | |
\begin{proposition}[Uniqueness of products]\label{produnique} | |
Any two products of the collection $\left\{X_i\right\}$ in $\mathcal{C}$ are | |
isomorphic by a unique isomorphism commuting with the projections. | |
\end{proposition} | |
This is a special case of a general ``abstract nonsense'' type result that we | |
shall see many more of in the sequel. | |
The precise statement is the following: let $X$ be a product of the | |
$\left\{X_i\right\}$ with projections $p_i : X \to X_i$, and let $Y$ be a | |
product of them too, with projections $q_i: Y \to X_i$. | |
Then the claim is that there is a \emph{unique} isomorphism | |
\[ f: X \to Y \] | |
such that the diagrams below commute for each $i \in I$: | |
\begin{equation} \label{prodcommutative} \xymatrix{ | |
X \ar[rd]^{p_i} \ar[rr]^f & & Y \ar[ld]_{q_i} \\ | |
& X_i. | |
}\end{equation} | |
\begin{proof} | |
This is a ``trivial'' result, and is part of a general fact that objects | |
with the same universal property are always canonically isomorphic. Indeed, note that the projections $p_i: X \to | |
X_i$ and the fact that mapping into $Y$ is the same as mapping into all the | |
$X_i$ gives a unique map $f: X \to Y$ making the diagrams | |
\eqref{prodcommutative} commute. The same reasoning (applied to the $q_i: Y \to | |
X_i$) gives a map $g: Y \to X$ making the diagrams | |
\begin{equation} \label{prodcommutative2} \xymatrix{ | |
Y \ar[rd]^{q_i} \ar[rr]^g & & X \ar[ld]_{p_i} \\ | |
& X_i | |
}\end{equation} | |
commute. By piecing the two diagrams together, it follows that the composite $g \circ f$ makes the diagram | |
\begin{equation} \label{prodcommutative3} \xymatrix{ | |
X \ar[rd]^{p_i} \ar[rr]^{g \circ f} & & X \ar[ld]_{p_i} \\ | |
& X_i | |
}\end{equation} | |
commute. | |
But the identity $1_X: X \to X$ also would make \eqref{prodcommutative3} | |
commute, and the \emph{uniqueness} assertion in the definition of the product | |
shows that $g \circ f = 1_X$. Similarly, $f \circ g = 1_Y$. We are done. | |
\end{proof} | |
\begin{remark} | |
If we reverse the arrows in the above construction, | |
the universal property obtained (known as the ``coproduct'') characterizes | |
disjoint unions in the category of sets and free products in the category of | |
groups. | |
That is, to map \emph{out} of a coproduct of objects $\left\{X_i\right\}$ is the same as | |
mapping out of each of these. We shall later study this construction more | |
generally. | |
\end{remark} | |
\begin{exercise} | |
Let $P$ be a poset, and make $P$ into a category as in \cref{posetcategory}. | |
Fix $x, y \in P$. Show that the \emph{product} of $x,y$ is the greatest lower | |
bound of $\left\{x,y\right\}$ (if it exists). This claim holds more generally | |
for arbitrary subsets of $P$. | |
In particular, consider the poset of subsets of a given set $S$. Then the | |
``product'' in this category corresponds to the intersection of subsets. | |
\end{exercise} | |
We shall, in this section, investigate this notion of ``universality'' | |
more thoroughly. | |
\subsection{Initial and terminal objects} | |
We now introduce another example of universality, which is simpler but more | |
abstract than the products introduced in the previous section. | |
\begin{definition} | |
Let $\mathcal{C}$ be a category. An \textbf{initial object} in $\mathcal{C}$ is an | |
object $X \in \mathcal{C}$ with the property that $\hom_{\mathcal{C}}(X, Y)$ has one | |
element for all $Y \in \mathcal{C}$. | |
\end{definition} | |
So there is a unique map out of $X$ into each $Y \in \mathcal{C}$. | |
Note that this idea is faithful to the categorical spirit of describing objects | |
in terms of their mapping properties. Initial objects are very easy to map | |
\emph{out} of. | |
\begin{example} | |
If $\mathcal{C}$ is $\mathbf{Sets}$, then the empty set $\emptyset$ is an | |
initial object. There is a unique map from the empty set into any other set; | |
one has to make no decisions about where elements are to map when | |
constructing a map $\emptyset \to X$. | |
\end{example} | |
\begin{example} | |
In the category $\mathbf{Grp}$ of groups, the group consisting of one element | |
is an initial object. | |
\end{example} | |
Note that the initial object in $\mathbf{Grp}$ is \emph{not} that in | |
$\mathbf{Sets}$. This should not be too surprising, because $\emptyset$ cannot | |
be a group. | |
\begin{example} | |
Let $P$ be a poset, and make it into a category as in \cref{posetcategory}. | |
Then it is easy to see that an initial object of $P$ is the smallest object in | |
$P$ (if it exists). Note that this is equivalently the product of all the | |
objects in $P$. In general, the initial object of a category is not the product | |
of all objects in $\mathcal{C}$ (this does not even make sense for a large | |
category). | |
\end{example} | |
There is a dual notion, called a \textit{terminal object}, where every object | |
can map into it in precisely one way. | |
\begin{definition} | |
A \textbf{terminal object} in a category $\mathcal{C}$ is an object $Y \in | |
\mathcal{C}$ such that $\hom_{\mathcal{C}}(X, Y) = \ast$ for each $X \in \mathcal{C}$. | |
\end{definition} | |
Note that an initial object in $\mathcal{C}$ is the same as a terminal object | |
in $\mathcal{C}^{op}$, and vice versa. As a result, it suffices to prove | |
results about initial objects, and the corresponding results for terminal | |
objects will follow formally. | |
But there is a fundamental difference between initial and terminal objects. | |
Initial objects are characterized by how one maps \emph{out of} them, while | |
terminal objects are characterized by how one maps \emph{into} them. | |
\begin{example} | |
The one point set is a terminal object in $\mathbf{Sets}$. | |
\end{example} | |
The important thing about the next ``theorems'' is the conceptual framework. | |
\begin{proposition}[Uniqueness of the initial (or terminal) object] | |
\label{initialunique} | |
Any two initial (resp. terminal) objects in $\mathcal{C}$ are isomorphic by a | |
unique isomorphism. | |
\end{proposition} | |
\begin{proof} | |
The proof is easy. We do it for terminal objects. Say $Y, Y'$ are | |
terminal objects. Then $\hom(Y, Y')$ and $\hom(Y', Y)$ are one | |
point sets. So there are unique maps $f: Y \to Y', g: Y' \to Y$, whose composites | |
must be the identities: we know that $\hom(Y, Y) , \hom(Y', Y')$ | |
are one-point sets, so the composites have no other choice to be the | |
identities. This means that the maps $f: Y \to Y', g: Y' \to Y$ are | |
isomorphisms. | |
\end{proof} | |
There is a philosophical point to be made here. We have characterized an object | |
uniquely in terms of mapping properties. We have characterized it | |
\emph{uniquely up to unique isomorphism,} which is really the best one can do | |
in mathematics. Two sets are not generally the ``same,'' but they may be | |
isomorphic up to unique isomorphism. They are different, | |
but the sets are isomorphic up | |
to unique isomorphism. | |
Note also that the argument was essentially similar to that of \cref{produnique}. | |
In fact, we could interpret \cref{produnique} as a special case of | |
\cref{initialunique}. | |
If $\mathcal{C}$ is a category and $\left\{X_i\right\}_{i \in I}$ is a family | |
of objects in $\mathcal{C}$, then we can define a category $\mathcal{D}$ as | |
follows. An object of $\mathcal{D}$ is the data of an object $Y \in | |
\mathcal{C}$ and morphisms $f_i: Y \to X_i$ for all $i \in I$. | |
A morphism between objects $(Y, \left\{f_i: Y \to X_i\right\})$ and $(Z, | |
\left\{g_i: Z \to X_i\right\})$ is | |
a map $Y \to Z$ making the obvious diagrams commute. Then a product $\prod X_i$ | |
in $\mathcal{C}$ is the same thing as a terminal object in $\mathcal{D}$, as | |
one easily checks from the definitions. | |
\subsection{Push-outs and pull-backs} | |
Let $\mathcal{C}$ be a category. | |
Now we are going to talk about more examples of universal constructions, which can all be | |
phrased via initial or terminal objects in some category. This, | |
therefore, is the proof for the uniqueness up to unique | |
isomorphism of \emph{everything} we will do in this | |
section. Later we will present these in more generality. | |
Suppose we have objects $A, B, C, X \in \mathcal{C}$. | |
\begin{definition} | |
A commutative square | |
\[ | |
\xymatrix{ | |
A \ar[d] \ar[r] & B \ar[d] \\ | |
C \ar[r] & X}. | |
\] | |
is a \textbf{pushout square} (and $X$ is called the \textbf{push-out}) if, | |
given a commutative diagram | |
\[ \xymatrix{ | |
A \ar[r] \ar[d] & B \ar[d] \\ | |
C \ar[r] & Y \\ | |
}\] | |
there is a unique map $X \to Y$ making the following diagram commute: | |
\[ | |
\xymatrix{ | |
A \ar[d] \ar[r] & B \ar[d] \ar[rdd] \\ | |
C \ar[r] \ar[rrd] & X \ar[rd] \\ | |
& & Y'} | |
\] | |
Sometimes push-outs are also called \textbf{fibered coproducts}. | |
We shall also write $X = C \sqcup_A B$. | |
\end{definition} | |
In other words, to map out of $X = C \sqcup_A B$ into some object $Y$ is to | |
give maps $B \to Y, C \to Y$ whose restrictions to $A$ are the same. | |
The next few examples will rely on notions to be introduced later. | |
\begin{example} | |
The following is a pushout square in the category of abelian groups: | |
\[ \xymatrix{ | |
\mathbb{Z}/2 \ar[r] \ar[d] & \mathbb{Z}/4 \ar[d] \\ | |
\mathbb{Z}/6 \ar[r] & \mathbb{Z}/12 | |
}\] | |
In the category of groups, the push-out is actually | |
$\mathrm{SL}_2(\mathbb{Z})$, though we do not prove it. The point is that | |
the property of a square's being a | |
push-out is actually dependent on the category. | |
In general, to construct a push-out of groups $C \sqcup_A B$, one constructs | |
the direct sum $C \oplus B$ and quotients by the subgroup generated by | |
$(a, a)$ (where $a \in A$ is identified with its image in $C \oplus B$). | |
We shall discuss this later, more thoroughly, for modules over a ring. | |
\end{example} | |
\begin{example} | |
Let $R$ be a commutative ring and let $S$ and $Q$ be two commutative | |
$R$-algebras. In other words, suppose | |
we have two maps of rings $s:R\rightarrow S$ and $q:R\rightarrow Q$. Then we | |
can fit this information together | |
into a pushout square: | |
\[ \xymatrix{ | |
R \ar[r] \ar[d] & S \ar[d] \\ | |
Q \ar[r] &X | |
}\] | |
It turns out that the pushout in this case is the tensor product of algebras | |
$S\otimes_R Q$ (see \cref{tensprodalg} for the construction). This is particularly important | |
in algebraic geometry as the dual construction will give the correct notion of | |
``products'' in the category of ``schemes'' over | |
a field.\end{example} | |
\begin{proposition} | |
Let $\mathcal{C}$ be any category. | |
If the push-out of | |
the diagram | |
\[ \xymatrix{ | |
A \ar[d] \ar[r] & B \\ | |
C | |
}\] | |
exists, it is unique up to unique isomorphism. | |
\end{proposition} | |
\begin{proof} | |
We can prove this in two ways. One is to suppose that there were two pushout | |
squares: | |
\[ | |
\xymatrix{ | |
A \ar[d] \ar[r] & B \ar[d] \\ | |
C \ar[r] & X \\ | |
} | |
\quad \quad | |
\xymatrix{ | |
A \ar[d] \ar[r] & B \ar[d] \\ | |
C \ar[r] & X' \\ | |
} | |
\] | |
Then there are unique maps $f:X \to X', g: X' \to X$ from the universal property. | |
In detail, these maps fit into commutative diagrams | |
\[ | |
\xymatrix{ | |
A \ar[d] \ar[r] & B \ar[d] \ar[rdd] \\ | |
C \ar[r] \ar[rrd] & X \ar[rd]^f\\ | |
& & X' | |
} | |
\quad \quad | |
\xymatrix{ | |
A \ar[d] \ar[r] & B \ar[d] \ar[rdd] \\ | |
C \ar[r] \ar[rrd] & X' \ar[rd]^g\\ | |
& & X | |
} | |
\] | |
Then $g \circ f$ and $f \circ g$ are the identities of $X, X'$ again by | |
\emph{uniqueness} of the map in the definition of the push-out. | |
Alternatively, we can phrase push-outs in terms of initial objects. We could | |
consider the category of all diagrams as above, | |
\[ \xymatrix{ | |
A \ar[d] \ar[r] & B \ar[d] \\ | |
C \ar[r] & D | |
},\] | |
where $A \to B, A \to C$ are fixed and $D$ varies. | |
The morphisms in this category of diagrams consist of commutative | |
diagrams. Then the initial | |
object in this category is the push-out, as one easily checks. | |
\end{proof} | |
Often when studying categorical constructions, one can create a kind of | |
``dual''construction by reversing the direction of the arrows. This is exactly | |
the | |
relationship between the push-out construction and the pull-back | |
construction to be described below. | |
So suppose we have two morphisms $A \to C$ and $B\to C$, forming a diagram | |
\[ \xymatrix{ | |
& B \ar[d] \\ | |
A \ar[r] & C. | |
}\] | |
\begin{definition} | |
The \textbf{pull-back} or \textbf{fibered product} of the above | |
diagram is an object $P$ with two morphisms $P\to B$ and $P\to | |
C$ such that the following diagram commutes: | |
\[ \xymatrix { | |
P \ar[d] \ar[r] & B \ar[d]\\ | |
A\ar[r] & C }\] | |
Moreover, the object $P$ is required to be universal in the following sense: given any $P'$ | |
and maps $P'\to A$ and $P'\to B$ making the square commute, there is a | |
unique map | |
$P'\to P$ making the following diagram commute: | |
\[ | |
\xymatrix{ | |
P' \ar[rd] \ar[rrd] \ar[ddr] \\ | |
& P \ar[d] \ar[r] & B \ar[d] \\ | |
& A \ar[r] & C }\] | |
We shall also write $P = B \times_C A$. | |
\end{definition} | |
\begin{example} | |
In the category $\mathbf{Set}$ of sets, if we have sets $A, B, C$ with maps $f: | |
A \to C, g: B \to C$, then the fibered product $A \times_C B$ consists of | |
pairs $(a,b) \in A \times B$ such that $f(a) = g(b)$. | |
\end{example} | |
\begin{example}[Requires prerequisites not developed yet] The next example may | |
be omitted without loss of continuity. | |
As said above, the fact that the tensor product of algebras is | |
a push-out in the category of | |
commutative $R$-algebras allows for the correct notion of the ``product'' of | |
schemes. We now elaborate on this example: naively one would think that we | |
could pick the underlying space of the product scheme to just be the topological | |
product of two Zariski topologies. However, it is an easy exercise to check | |
that the product of two Zariski topologies in general is not Zariski! This | |
motivates | |
the need for a different concept. | |
Suppose we have a field $k$ and two $k$-algebras $A$ and $B$ and let | |
$X=\spec(A)$and $Y=\spec(B)$ be the affine $k$-schemes corresponding to $A$ and | |
$B$. Consider the following pull-back diagram: | |
\[ | |
\xymatrix{ | |
X\times_{\spec(k)} Y \ar[d] \ar[r] &X \ar[d]\\ | |
Y \ar[r] &\spec(k) }\] | |
Now, since $\spec$ is a contravariant functor, the arrows in this pull-back | |
diagram have been flipped; so in fact, $X\times_{\spec(k)} Y$ is actually | |
$\spec(A\otimes _k B)$. This construction is motivated by the following example: | |
let $A=k[x]$ and $B=k[y]$. Then $\spec(A)$ and $\spec(B)$ are both affine lines | |
$\mathbb{A}^1_k$ so we want a suitable notion of product that makes the product | |
of $\spec(A)$ and $\spec(B)$ the affine plane. The pull-back construction is the | |
correct one since $\spec(A)\times_{\spec(k)} \spec(B)=\spec(A\otimes_k | |
B)=\spec(k[x,y])=\mathbb{A}^2_k$. | |
\end{example} | |
\subsection{Colimits} | |
We now want to generalize the push-out. | |
Instead of a shape with $A,B,C$, we do something more general. | |
Start with a small category $I$: recall that \emph{smallness} means that the objects of $I$ | |
form a set. $I$ is to be called the \textbf{indexing | |
category}. One is supposed to picture | |
is that $I$ is something like the category | |
\[ | |
\xymatrix{ | |
\ast \ar[d] \ar[r] & \ast \\ | |
\ast | |
} | |
\] | |
or the category | |
\[ \ast \rightrightarrows \ast. \] | |
We will formulate the notion of a \textbf{colimit} which will specialize to the | |
push-out when $I$ is the first case. | |
So we will look at functors | |
\[ F: I \to \mathcal{C}, \] | |
which in the case of the three-element category, will just | |
correspond to | |
diagrams | |
\[ \xymatrix{A \ar[d] \ar[r] & B \\ C}. \] | |
We will call a \textbf{cone} on $F$ (this is an ambiguous term) an object $X | |
\in \mathcal{C}$ equipped with maps $F_i \to X, \forall i \in I$ such that for | |
all maps $i \to | |
i' \in I$, the diagram below commutes: | |
\[ \xymatrix{ | |
F_i \ar[d] \ar[r] & X \\ | |
F_{i'} \ar[ru] | |
}.\] | |
An example would be a cone on the three-element category above: then | |
this is just a commutative diagram | |
\[ \xymatrix{ | |
A \ar[r]\ar[d] & B \ar[d] \\ | |
C \ar[r] & D | |
}.\] | |
\newcommand{\colim}{\mathrm{colim}} | |
\begin{definition} | |
The \textbf{colimit} of the diagram $F: I \to \mathcal{C}$, written as $\colim | |
F$ or $\colim_I F $ or $\varinjlim_I F$, if it exists, is a cone $F \to X$ with | |
the property that if $F \to Y$ is any other cone, then there is a unique map $X | |
\to Y$ making the diagram | |
\[ \xymatrix{ | |
F \ar[rd] \ar[r] & X \ar[d] \\ | |
& Y | |
}\] | |
commute. (This means that the corresponding diagram with $F_i$ replacing $F$ | |
commutes for each $i \in I$.) | |
\end{definition} | |
We could form a category $\mathcal{D}$ where the objects are the cones $F \to | |
X$, and the morphisms from $F \to X$ and $F \to Y$ are the maps $X \to Y$ that | |
make all the obvious diagrams commute. In this case, it is easy to see that a | |
\emph{colimit} of the diagram is just an initial object in $\mathcal{D}$. | |
In any case, we see: | |
\begin{proposition} | |
$\colim F$, if it exists, is unique up to unique isomorphism. | |
\end{proposition} | |
Let us go through some examples. We already looked at push-outs. | |
\begin{example} | |
Consider the category $I$ visualized as | |
\[ \ast, \ast, \ast, \ast. \] | |
So $I$ consists of four objects with no non-identity morphisms. | |
A functor $F: I \to \mathbf{Sets}$ is just a list of four sets $A, B, C, D$. | |
The colimit is just the disjoint union $A \sqcup B \sqcup C \sqcup D$. This is | |
the universal property of the disjoint union. To map out of the disjoint union | |
is the same thing as mapping out of each piece. | |
\end{example} | |
\begin{example} | |
Suppose we had the same category $I$ but the functor $F$ took values in the | |
category of abelian groups. Then $F$ | |
corresponds, again, to a list of four abelian groups. The colimit is the direct | |
sum. Again, the direct sum is characterized by the same universal property. | |
\end{example} | |
\begin{example} | |
Suppose we had the same $I$ ($\ast, \ast, \ast, \ast$) the functor took its | |
value in the category of groups. Then the colimit is the | |
free product of the four groups. | |
\end{example} | |
\begin{example} | |
Suppose we had the same $I$ and the category $\mathcal{C}$ was of commutative | |
rings with unit. Then the colimit is the tensor product. | |
\end{example} | |
So the idea of a colimit unifies a whole bunch of constructions. | |
Now let us take a different example. | |
\begin{example} | |
Take | |
\[ I = \ast \rightrightarrows \ast. \] | |
So a functor $I \to \mathbf{Sets}$ is a diagram | |
\[ A \rightrightarrows B. \] | |
Call the two maps $f,g: A \to B$. To get the colimit, we take $B$ and mod out | |
by the equivalence relation generated by $f(a) \sim g(a)$. | |
To hom out of this is the same thing as homming out of $B$ such that the | |
pullbacks to $A$ are the same. | |
This is the relation \textbf{generated} as above, not just as above. It can get | |
tricky. | |
\end{example} | |
\begin{definition} | |
When $I$ is just a bunch of points $\ast, \ast, \ast, \dots$ with no | |
non-identity morphisms, then the | |
colimit over $I$ is called the \textbf{coproduct}. | |
\end{definition} | |
We use the coproduct to mean things like direct sums, disjoint unions, and | |
tensor products. | |
If $\left\{A_i, i \in I\right\}$ is a collection of objects in some category, | |
then we find the universal property of the coproduct can be stated succinctly: | |
\[ \hom_{\mathcal{C}}(\bigsqcup_I A_i, B) = \prod \hom_{\mathcal{C}}(A_i, B). \] | |
\begin{definition} | |
When $I$ is $\ast \rightrightarrows \ast$, the colimit is called the | |
\textbf{coequalizer}. | |
\end{definition} | |
\begin{theorem} \label{coprodcoequalsufficeforcocomplete} | |
If $\mathcal{C}$ has all coproducts and coequalizers, then it has all colimits. | |
\end{theorem} | |
\begin{proof} | |
Let $F: I \to \mathcal{C}$ be a functor, where $I$ is a small category. We | |
need to obtain an object $X$ with morphisms | |
\[ Fi \to X, \quad i \in I \] | |
such that for each $f: i \to i'$, the diagram below commutes: | |
\[ | |
\xymatrix{ | |
Fi \ar[d] \ar[r] & Fi' \ar[ld] \\ | |
X | |
} | |
\] | |
and such that $X$ is universal among such diagrams. | |
To give such a diagram, however, is equivalent to giving a collection of maps | |
\[ Fi \to X \] | |
that satisfy some conditions. So $X$ should be thought of as a quotient of the | |
coproduct $\sqcup_i Fi$. | |
Let us consider the coproduct $\sqcup_{i \in I, f} Fi$, where $f$ ranges over | |
all | |
morphisms in the category $I$ that start from $i$. | |
We construct two maps | |
\[ \sqcup_f Fi \rightrightarrows \sqcup_f Fi, \] | |
whose coequalizer will be that of $F$. The first map is the identity. The | |
second map sends a factor | |
\end{proof} | |
\subsection{Limits} | |
As in the example with pull-backs and push-outs and products and coproducts, | |
one can define a limit by using the exact same universal property above | |
just with | |
all the arrows reversed. | |
\begin{example} The product is an example of a limit where the indexing | |
category is a small category $I$ with no morphisms other than the identity. This | |
example | |
shows the power of universal constructions; by looking at colimits and limits, | |
a whole variety of seemingly unrelated mathematical constructions are shown | |
to be | |
in the same spirit. | |
\end{example} | |
\subsection{Filtered colimits} | |
\emph{Filtered colimits} are colimits | |
over special indexing categories $I$ which look like totally ordered sets. | |
These have several convenient properties as compared to general colimits. | |
For instance, in the category of \emph{modules} over a ring (to be studied in | |
\rref{foundations}), we shall see that filtered colimits actually | |
preserve injections and surjections. In fact, they are \emph{exact.} This is | |
not true in more general categories which are similarly structured. | |
\begin{definition} | |
An indexing category is \textbf{filtered} if the following hold: | |
\begin{enumerate} | |
\item Given $i_0, i_1 \in I$, there is a third object $i \in I$ such that both | |
$i_0, i_1$ map into $i$. | |
So there is a diagram | |
\[ \xymatrix{ | |
i_0 \ar[rd] \\ | |
& i \\ | |
i_1 \ar[ru] | |
}.\] | |
\item Given any two maps $i_0 \rightrightarrows i_1$, there exists $i$ and $i_1 | |
\to i$ such that the two maps $i_0 \rightrightarrows i$ are equal: | |
intuitively, any two ways | |
of pushing an object into another can be made into the same eventually. | |
\end{enumerate} | |
\end{definition} | |
\begin{example} | |
If $I$ is the category | |
\[ \ast \to \ast \to \ast \to \dots, \] | |
i.e. the category generated by the poset $\mathbb{Z}_{\geq 0}$, then that is | |
filtered. | |
\end{example} | |
\begin{example} | |
If $G$ is a torsion-free abelian group, the category $I$ of finitely generated | |
subgroups of $G$ and inclusion maps is filtered. We don't actually need the | |
lack of torsion. | |
\end{example} | |
\begin{definition} | |
Colimits over a filtered category are called \textbf{filtered colimits}. | |
\end{definition} | |
\begin{example} | |
Any torsion-free abelian group is the filtered colimit of its finitely | |
generated subgroups, which are free abelian groups. | |
\end{example} | |
This gives a simple approach for showing that a torsion-free abelian group is | |
flat. | |
\begin{proposition} | |
If $I$ is filtered\footnote{Some people say filtering.} and $\mathcal{C} = | |
\mathbf{Sets}, \mathbf{Abgrp}, \mathbf{Grps}$, etc., and $F: I \to \mathcal{C}$ | |
is a functor, then $\colim_I F$ exists and is given by the disjoint union of | |
$F_i, i \in I$ modulo the relation $x \in F_i$ is equivalent to $x' \in F_{i'}$ | |
if $x$ maps to $x'$ under $F_i \to F_{i'}$. This is already an equivalence | |
relation. | |
\end{proposition} | |
The fact that the relation given above is transitive uses the filtering of the | |
indexing set. Otherwise, we would need to use the relation generated by it. | |
\begin{example} | |
Take $\mathbb{Q}$. This is the filtered colimit of the free submodules | |
$\mathbb{Z}(1/n)$. | |
Alternatively, choose a sequence of numbers $m_1 , m_2, \dots, $ such that for | |
all $p, n$, we have $p^n \mid m_i$ for $i \gg 0$. Then we have a sequence of | |
maps | |
\[ \mathbb{Z} \stackrel{m_1}{\to} \mathbb{Z} \stackrel{m_2}{\to}\mathbb{Z} | |
\to \dots. \] | |
The colimit of this is $\mathbb{Q}$. There is a quick way of seeing this, which | |
is left to the reader. | |
\end{example} | |
When we have a functor $F: I \to \mathbf{Sets}, \mathbf{Grps}, | |
\mathbf{Modules}$ taking values in a ``nice'' category (e.g. the category of | |
sets, modules, etc.), one can construct the colimit by taking the union of the | |
$F_i, i \in I$ and quotienting by the equivalence relation $x \in F_i \sim x' | |
\in F_{i'}$ if $f: i \to i'$ sends $x$ into $x'$. This is already an | |
equivalence relation, as one can check. | |
Another way of saying this is that we have the disjoint union of the $F_i$ | |
modulo the relation that $a \in F_i$ and $b \in F_{i'}$ are equivalent if and | |
only if there is a later $i''$ with maps $i \to i'', i' \to i''$ such that | |
$a,b$ both map to the same thing in $F_{i''}$. | |
One of the key properties of filtered colimits is that, in ``nice'' categories they commute with | |
finite limits. | |
\begin{proposition} | |
In the category of sets, filtered colimits and finite limits commute with each | |
other. | |
\end{proposition} | |
The reason this result is so important is that, as we shall see, it will imply | |
that in categories such as the category of $R$-modules, filtered colimits | |
preserve \emph{exactness}. | |
\begin{proof} | |
Let us show that filtered colimits commute with (finite) products in the | |
category of sets. The case of an equalizer is similar, and finite limits can be | |
generated from products and equalizers. | |
So let $I$ be a filtered category, and $\left\{A_i\right\}_{i \in I}, | |
\left\{B_i\right\}_{i \in I}$ | |
be functors from $I \to \mathbf{Sets}$. | |
We want to show that | |
\[ \varinjlim_I (A_i \times B_i) = \varinjlim_I A_i \times \varinjlim_I B_i . \] | |
To do this, note first that there is a map in the direction $\to$ because of | |
the natural maps $\varinjlim_I (A_i \times B_i) \to \varinjlim_I A_i$ and | |
$\varinjlim_I (A_i \times B_i) \to \varinjlim_I B_i$. | |
We want to show that this is an isomorphism. | |
Now we can write the left side as the disjoint union $\bigsqcup_I (A_i \times | |
B_i)$ modulo the equivalence relation that $(a_i, b_i)$ is related to $(a_j, | |
b_j)$ if there exist morphisms $i \to k, j \to k$ sending $(a_i, b_i), (a_j, | |
b_j)$ to the same object in $A_k \times B_k$. | |
For the left side, we have to work with pairs: that is, an element of | |
$\varinjlim_I A_i \times \varinjlim_I B_i$ consists of a pair $(a_{i_1}, | |
b_{i_2})$ | |
with two pairs $(a_{i_1}, b_{i_2}), (a_{j_1}, b_{j_2})$ equivalent if there exist | |
morphisms $i_1,j_1 \to k_1 $ and $i_2, j_2 \to k_2$ such that both have the | |
same image in $A_{k_1} \times A_{k_2}$. It is easy to see that these amount to | |
the same thing, because of the filtering condition: we can always modify an | |
element of $A_{i} \times B_{j}$ to some $A_{k} \times B_k$ for $k$ receiving | |
maps from $i, j$. | |
\end{proof} | |
\begin{exercise} | |
Let $A$ be an abelian group, $e: A \to A$ an \emph{idempotent} operator, i.e. | |
one such that $e^2 = e$. Show that $eA$ can be obtained as the filtered colimit | |
of | |
\[ A \stackrel{e}{\to} A \stackrel{e}{\to} A \dots. \] | |
\end{exercise} | |
\subsection{The initial object theorem} | |
We now prove a fairly nontrivial result, due to Freyd. This gives a sufficient | |
condition for the existence of initial objects. | |
We shall use it in proving the adjoint functor theorem below. | |
Let $\mathcal{C}$ be a category. Then we recall that $A \in \mathcal{C}$ if | |
for each $X \in \mathcal{C}$, there is a \emph{unique} $A \to X$. | |
Let us consider the weaker condition that for each $ X \in \mathcal{C}$, there | |
exists \emph{a} map $A \to X$. | |
\begin{definition} Suppose $\mathcal{C}$ has equalizers. | |
If $A \in \mathcal{C}$ is such that $\hom_{\mathcal{C}}(A, X) \neq \emptyset$ | |
for each $X \in \mathcal{C}$, then $X$ is called \textbf{weakly initial.} | |
\end{definition} | |
We now want to get an initial object from a weakly initial object. | |
To do this, note first that if $A$ is weakly initial and $B$ is any object | |
with a morphism $B \to A$, then $B$ is weakly initial too. So we are going to | |
take | |
our initial object to be a very small subobject of $A$. | |
It is going to be so small as to guarantee the uniqueness condition of an | |
initial object. To make it small, we equalize all endomorphisms. | |
\begin{proposition} \label{weakinitial} | |
If $A$ is a weakly initial object in $\mathcal{C}$, | |
then the equalizer of all endomorphisms $A \to A$ is initial for $\mathcal{C}$. | |
\end{proposition} | |
\begin{proof} | |
Let $A'$ be this equalizer; it is endowed with a morphism $A'\to A$. Then let | |
us recall what this means. For any two | |
endomorphisms $A \rightrightarrows A$, the two pull-backs $A' | |
\rightrightarrows A$ are equal. Moreover, if $B \to A$ is a morphism that has | |
this property, then $B$ factors uniquely through $A'$. | |
Now $A' \to A$ is a morphism, so by the remarks above, $A'$ is weakly initial: | |
to each $X \in \mathcal{C}$, there exists a morphism $A' \to X$. | |
However, we need to show that it is unique. | |
So suppose given two maps $f,g: A' \rightrightarrows X$. We are going to show | |
that they are equal. If not, consider their equalizer $O$. | |
Then we have a morphism $O \to A'$ such that the post-compositions with $f,g$ | |
are equal. But by weak initialness, there is a map $A \to O$; thus we get a | |
composite | |
\[ A \to O \to A'. \] | |
We claim that this is a \emph{section} of the embedding $A'\to A$. | |
This will prove the result. Indeed, we will have constructed a section $A \to | |
A'$, and since it factors through $O$, the two maps | |
\[ A \to O \to A' \rightrightarrows X \] | |
are equal. Thus, composing each of these with the inclusion $A' \to A$ shows | |
that $f,g$ were equal in the first place. | |
Thus we are reduced to proving: | |
\begin{lemma} | |
Let $A$ be an object of a category $\mathcal{C}$. Let $A'$ be the equalizer of | |
all endomorphisms of $A$. Then any morphism $A \to A'$ is a section of the | |
inclusion $A' \to A$. | |
\end{lemma} | |
\begin{proof} | |
Consider the canonical inclusion $i: A' \to A$. We are given some map $s: A | |
\to A'$; we must show that $si = 1_{A'}$. | |
Indeed, consider the composition | |
\[ A' \stackrel{i}{\to} A \stackrel{s}{\to} A' \stackrel{i}{\to} A .\] | |
Now $i$ equalizes endomorphisms of $A$; in particular, this composition is the | |
same as | |
\[ A' \stackrel{i}{\to} A \stackrel{\mathrm{id}}{\to} A; \] | |
that is, it equals $i$. So the map $si: A' \to A$ has the property that $isi = | |
i$ as maps $A' \to A$. But $i$ being a monomorphism, it follows that $si = | |
1_{A'}$. | |
\end{proof} | |
\end{proof} | |
\begin{theorem}[Freyd] \label{initialobjectthm} | |
Let $\mathcal{C}$ be a category admitting all small limits.\footnote{We shall | |
later call such a category \textbf{complete}.} Then $\mathcal{C}$ has an initial | |
object if and only if the following \textbf{solution set condition holds:} | |
there is a set $\left\{X_i, i \in I\right\}$ of objects in $\mathcal{C}$ such | |
that any $X \in \mathcal{C}$ can be mapped into by one of these. | |
\end{theorem} | |
The idea is that the family $\left\{X_i\right\}$ is somehow weakly universal | |
\emph{together.} | |
\begin{proof} | |
If $\mathcal{C}$ has an initial object, we may just consider that as the | |
family $\left\{X_i\right\}$: we can hom out (uniquely!) from a universal | |
object into anything, or in other words a universal object is weakly universal. | |
Suppose we have a ``weakly universal family'' $\left\{X_i\right\}$. Then the | |
product $\prod X_i$ is weakly universal. Indeed, if $X \in \mathcal{C}$, | |
choose some $i'$ and a morphism $X_{i'} \to X$ by the hypothesis. Then this map | |
composed with the projection from the product gives a map $\prod X_i \to | |
X_{i'} \to X$. | |
\cref{weakinitial} now implies that $\mathcal{C}$ has an initial object. | |
\end{proof} | |
\subsection{Completeness and cocompleteness} | |
\begin{definition}\label{completecat} A category $\mathcal{C}$ is said to be \textbf{complete} if for every | |
functor $F:I\rightarrow \mathcal{C}$ where $I$ is a small category, the limit | |
$\lim F$ exists (i.e. $\mathcal{C}$ has all small limits). If all colimits exist, then $\mathcal{C}$ is said to be | |
\textbf{cocomplete}. | |
\end{definition} | |
If a category is complete, various nice properties hold. | |
\begin{proposition} If $\mathcal{C}$ is a complete category, the following | |
conditions are true: | |
\begin{enumerate} | |
\item{all (finite) products exist} | |
\item{all pull-backs exist} | |
\item{there is a terminal object} | |
\end{enumerate} | |
\end{proposition} | |
\begin{proof} The proof of the first two properties is trivial since they can | |
all be expressed as limits; for the proof of the existence of a terminal | |
object, consider the empty diagram $F:\emptyset \rightarrow \mathcal{C}$. Then | |
the | |
terminal object is just $\lim F$. | |
\end{proof} | |
Of course, if one dualizes everything we get a theorem about cocomplete | |
categories which is proved in essentially the same manner. More is true | |
however; it turns out that finite (co)completeness are equivalent to the | |
properties above if one requires the finiteness condition for the existence of | |
(co)products. | |
\subsection{Continuous and cocontinuous functors} | |
\subsection{Monomorphisms and epimorphisms} | |
We now wish to characterize monomorphisms and epimorphisms in a purely | |
categorical setting. In categories where there is an underlying set the notions | |
of injectivity and surjectivity makes sense but in category theory, one | |
does not | |
in a sense have ``access'' to the internal structure of objects. In this light, | |
we make the following definition. | |
\begin{definition} | |
A morphism $f:X \to Y$ is a \textbf{monomorphism} if for any two morphisms | |
$g_1:X'\rightarrow X$ and $g_2:X'\rightarrow X$, we have that $f g_1 = f g_2$ | |
implies $g_1=g_2$. A morphism $f:X\rightarrow Y$ is an \textbf{epimorphism} if for any two | |
maps $g_1:Y\rightarrow Y'$ and $g_2:Y\rightarrow Y'$, we have that $g_1 f = g_2 | |
f$ implies $g_1 = g_2$. | |
\end{definition} | |
So $f: X \to Y$ is a monomorphism if whenever $X'$ is another object in | |
$\mathcal{C}$, the map | |
\[ \hom_{\mathcal{C}}(X', X) \to \hom_{\mathcal{C}}(X', Y) \] | |
is an injection (of sets). Epimorphisms in a category are defined similarly; | |
note that neither definition makes any reference to \emph{surjections} of sets. | |
The reader can easily check: | |
\begin{proposition} \label{compositeofmono} | |
The composite of two monomorphisms is a monomorphism, as is the composite of | |
two epimorphisms. | |
\end{proposition} | |
\begin{exercise} | |
Prove \cref{compositeofmono}. | |
\end{exercise} | |
\begin{exercise} | |
The notion of ``monomorphism'' can be detected using only the notions of | |
fibered product and isomorphism. To see this, suppose $i: X \to Y$ is a | |
monomorphism. Show that the diagonal | |
\[ X \to X \times_Y X \] | |
is an isomorphism. (The diagonal map is such that the two | |
projections to $X$ both give the identity.) Conversely, show that if $i: X \to Y$ is any morphism such | |
that the above diagonal map is an isomorphism, then $i$ is a monomorphism. | |
Deduce the following consequence: if $F: \mathcal{C} \to \mathcal{D}$ is a | |
functor that commutes with fibered products, then $F $ takes monomorphisms to | |
monomorphisms. | |
\end{exercise} | |
\section{Yoneda's lemma} | |
\add{this section is barely fleshed out} | |
Let $\mathcal{C}$ be a category. | |
In general, we have said that there is no way to study an object in a | |
category other than by considering maps into and out of it. | |
We will see that essentially everything about $X \in \mathcal{C}$ can be | |
recovered from these hom-sets. | |
We will thus get an embedding of $\mathcal{C}$ into a category of functors. | |
\subsection{The functors $h_X$} | |
We now use the structure of a category to construct hom functors. | |
\begin{definition} | |
Let $X \in \mathcal{C}$. We define the contravariant functor $h_X: \mathcal{C} | |
\to \mathbf{Sets}$ via | |
\[ h_X(Y) = \hom_{\mathcal{C}}(Y, X). \] | |
\end{definition} | |
This is, indeed, a functor. If $g: Y \to Y'$, then precomposition gives a map | |
of sets | |
\[ h_X(Y') \to h_X(Y), \quad f \mapsto f \circ g \] | |
which satisfies all the usual identities. | |
As a functor, $h_X$ encodes \emph{all} the information about | |
how one can map into $X$. | |
It turns out that one can basically recover $X$ from $h_X$, though. | |
\subsection{The Yoneda lemma} | |
Let $X \stackrel{f}{\to} X'$ be a morphism in $\mathcal{C}$. | |
Then for each $Y \in \mathcal{C}$, composition gives a map | |
\[ \hom_{\mathcal{C}}(Y, X) \to \hom_{\mathcal{C}}(Y, X'). \] | |
It is easy to see that this induces a \emph{natural} transformation | |
\[ h_{X} \to h_{X'}. \] | |
Thus we get a map of sets | |
\[ \hom_{\mathcal{C}}(X, X') \to \hom(h_X, h_{X'}), \] | |
where $h_X, h_{X'}$ lie in the category of contravariant functors $\mathcal{C} | |
\to \mathbf{Sets}$. | |
In other words, we have defined a \emph{covariant functor} | |
\[ \mathcal{C} \to \mathbf{Fun}(\mathcal{C}^{op}, \mathbf{Sets}). \] | |
This is called the \emph{Yoneda embedding.} The next result states that the | |
embedding is fully faithful. | |
\begin{theorem}[Yoneda's lemma] | |
\label{yonedalemma} | |
If $X, X' \in \mathcal{C}$, then the map | |
$\hom_{\mathcal{C}}(X, X') \to \hom(h_X, h_{X'})$ is a bijection. That is, | |
every natural transformation $h_X \to h_{X'}$ arises in one and only one way | |
from a morphism $X \to X'$. | |
\end{theorem} | |
\begin{theorem}[Strong Yoneda lemma] | |
\end{theorem} | |
\subsection{Representable functors} | |
We use the same notation of the preceding section: for a category | |
$\mathcal{C}$ and $X \in \mathcal{C}$, we let $h_X$ be the contravariant | |
functor $\mathcal{C} \to \mathbf{Sets}$ given by $Y \mapsto | |
\hom_{\mathcal{C}}(Y, X)$. | |
\begin{definition} | |
A contravariant functor $F: \mathcal{C} \to \mathbf{Sets}$ is | |
\textbf{representable} if it is naturally isomorphic to some $h_X$. | |
\end{definition} | |
The point of a representable functor is that it can be realized as maps into a | |
specific object. | |
In fact, let us look at a specific feature of the functor $h_X$. | |
Consider the object $\alpha \in h_X(X)$ that corresponds to the identity. | |
Then any morphism | |
\[ Y \to X \] | |
factors \emph{uniquely} | |
as \[ Y \to X \stackrel{\alpha}{\to } X \] | |
(this is completely trivial!) so that | |
any element of $h_X(Y)$ is a $f^*(\alpha)$ for precisely one $f: Y \to X$. | |
\begin{definition} | |
Let $F: \mathcal{C} \to \mathbf{Sets}$ be a contravariant functor. A | |
\textbf{universal object} for $\mathcal{C}$ is a pair $(X, \alpha)$ where $X | |
\in \mathcal{C}, \alpha \in F(X)$ such that the following condition holds: | |
if $Y$ is any object and $\beta \in F(Y)$, then there is a unique $f: Y \to X$ | |
such that $\alpha$ pulls back to $\beta$ under $f$. | |
In other words, $\beta = f^*(\alpha)$. | |
\end{definition} | |
So a functor has a universal object if and only if it is representable. | |
Indeed, we just say that the identity $X \to X$ is universal for $h_X$, and | |
conversely if $F$ has a universal object $(X, \alpha)$, then $F$ is naturally | |
isomorphic to $h_X$ (the isomorphism $h_X \simeq F$ being given by pulling | |
back $\alpha$ appropriately). | |
The article \cite{Vi08} by Vistoli contains a good introduction to and several | |
examples of this theory. | |
Here is one of them: | |
\begin{example} | |
Consider the contravariant functor $F: \mathbf{Sets} \to \mathbf{Sets}$ that | |
sends any set $S$ to its power set $2^S$ (i.e. the collection of subsets). | |
This is a contravariant functor: if $f: S \to T$, there is a morphism | |
\[ 2^T \to 2^S, \quad T' \mapsto f^{-1}(T'). \] | |
This is a representable functor. Indeed, the universal object can be taken as | |
the pair | |
\[ ( \left\{0,1\right\}, \left\{1\right\}). \] | |
To understand this, note that a subset $S;$ of $S$ determines its | |
\emph{characteristic function} $\chi_{S'}: S \to \left\{0,1\right\}$ that | |
takes the value $1$ on $S$ and $0$ elsewhere. | |
If we consider $\chi_{S'}$ as a morphism $ S \to \left\{0,1\right\}$, we see | |
that | |
\[ S' = \chi_{S'}^{-1}(\{1\}). \] | |
Moreover, the set of subsets is in natural bijection with the set of | |
characteristic functions, which in turn are precisely \emph{all} the maps $S | |
\to \left\{0,1\right\}$. From this the assertion is clear. | |
\end{example} | |
We shall meet some elementary criteria for the representability of | |
contravariant functors in the next subsection. For now, we note\footnote{The | |
reader unfamiliar with algebraic topology may omit these remarks.} that in | |
algebraic topology, one often works with the \emph{homotopy category} of | |
pointed CW complexes (where morphisms are pointed continuous maps modulo | |
homotopy), any contravariant functor that satisfies two relatively mild | |
conditions (a | |
Mayer-Vietoris condition and a condition on coproducts), is automatically | |
representable by a theorem of Brown. In particular, this implies that the | |
singular cohomology functors $H^n(-, G)$ (with coefficients in some group $G$) | |
are representable; the representing objects are the so-called | |
Eilenberg-MacLane spaces $K(G,n)$. See \cite{Ha02}. | |
\subsection{Limits as representable functors} | |
\add{} | |
\subsection{Criteria for representability} | |
Let $\mathcal{C}$ be a category. | |
We saw in the previous subsection that a representable functor must send | |
colimits to limits. | |
We shall now see that there is a converse under certain set-theoretic | |
conditions. | |
For simplicity, we start by stating the result for corepresentable functors. | |
\begin{theorem}[(Co)representability theorem] | |
Let $\mathcal{C}$ be a complete category, and let $F: \mathcal{C} \to | |
\mathbf{Sets}$ be a covariant functor. Suppose $F$ preserves limits and satisfies the solution set condition: | |
there is a set of objects $\left\{Y_\alpha\right\}$ such that, for any $X \in | |
\mathcal{C}$ and $x \in F(X)$, there is a morphism | |
\[ Y_\alpha \to X \] | |
carrying some element of $F(Y_\alpha)$ onto $x$. | |
Then $F$ is corepresentable. | |
\end{theorem} | |
\begin{proof} | |
To $F$, we associate the following \emph{category} $\mathcal{D}$. An object of | |
$\mathcal{D}$ is a pair $(x, X)$ where $x \in F(X)$ and $X \in \mathcal{C}$. | |
A morphism between $(x, X)$ and $(y, Y)$ is a map | |
\[ f:X \to Y \] | |
that sends $x$ into $y$ (via $F(f): F(X) \to F(Y)$). | |
It is easy to see that $F$ is corepresentable if and only if there is an initla | |
object in this category; this initial object is the ``universal object.'' | |
We shall apply the initial object theorem, \cref{initialobjectthm}. Let us first verify that | |
$\mathcal{D}$ is complete; this follows because $\mathcal{C}$ is and $F$ | |
preserves limits. So, for instance, the product of $(x, X)$ and $(y, Y)$ is | |
$((x,y), X \times Y)$; here $(x,y)$ is the element of $F(X) \times F(Y) = F(X | |
\times Y)$. | |
The solution set condition states that there is a weakly | |
initial family of objects, and the initial object theorem now implies that | |
there is an initial object. | |
\end{proof} | |
\section{Adjoint functors} | |
According to MacLane, ``Adjoint functors arise everywhere.'' We shall see | |
several examples of adjoint functors in this book (such as $\hom$ and the | |
tensor product). The fact that a functor has an adjoint often immediately | |
implies useful properties about it (for instance, that it commutes with either | |
limits or colimits); this will lead, for instance, to conceptual arguments | |
behind the right-exactness of the tensor product later on. | |
\subsection{Definition} | |
Suppose $\mathcal{C}, \mathcal{D}$ are categories, and let $F: \mathcal{C} \to | |
\mathcal{D}, G: \mathcal{D} \to \mathcal{C}$ be (covariant) functors. | |
\begin{definition} | |
$F, G$ are \textbf{adjoint functors} if there is a natural isomorphism | |
\[ \hom_{\mathcal{D}}(Fc, d) \simeq \hom_{\mathcal{C}}(c, Gd) \] | |
whenever $c \in \mathcal{C}, d \in \mathcal{D}$. $F$ is said to be the | |
\textbf{right adjoint,} and $G$ is the \textbf{left adjoint.} | |
\end{definition} | |
Here ``natural'' means that the two quantities are supposed to be considered | |
as functors $\mathcal{C}^{op} \times \mathcal{D} \to \mathbf{Set}$. | |
\begin{example} | |
There is a simple pair of adjoint functors between $\mathbf{Set}$ and $\mathbf{AbGrp}$. Here | |
$F$ sends a set $A$ to the free abelian group (see \cref{} for a discussion | |
of free modules over arbitrary rings) $\mathbb{Z}[A]$, while $G$ is | |
the ``forgetful'' functor that sends an abelian group to its underlying set. | |
Then $F$ and $G$ are adjoints. That is, to give a group-homomorphism | |
\[ \mathbb{Z}[A] \to G \] | |
for some abelian group $G$ | |
is the same as giving a map of \emph{sets} | |
\[ A \to G. \] | |
This is precisely the defining property of the free abelian group. | |
\end{example} | |
\begin{example} | |
In fact, most ``free'' constructions are just left adjoints. | |
For instance, recall the universal property of the free group $F(S)$ on a set $S$ (see | |
\cite{La02}): to give a group-homomorphism $F(S) \to G$ for $G$ any group is | |
the same as choosing an image in $G$ of each $s \in S$. | |
That is, | |
\[ \hom_{\mathbf{Grp}}(F(S), G) = \hom_{\mathbf{Sets}}(S, G). \] | |
This states that the free functor $S \mapsto F(S)$ is left adjoint to the | |
forgetful functor from $\mathbf{Grp}$ to $\mathbf{Sets}$. | |
\end{example} | |
\begin{example} | |
The abelianization functor $G \mapsto G^{ab} = G/[G, G]$ from $\mathbf{Grp} | |
\to \mathbf{AbGrp}$ is left adjoint to the | |
inclusion $\mathbf{AbGrp} \to \mathbf{Grp}$. | |
That is, if $G$ is a group and $A$ an abelian group, there is a natural | |
correspondence between homomorphisms $G \to A$ and $G^{ab} \to A$. | |
Note that $\mathbf{AbGrp}$ is a subcategory of $\mathbf{Grp}$ such that the | |
inclusion admits a left adjoint; in this situation, the subcategory is called | |
\textbf{reflective.} | |
\end{example} | |
\subsection{Adjunctions} | |
The fact that two functors are adjoint is encoded by a simple set of algebraic | |
data between them. | |
To see this, suppose $F: \mathcal{C} \to \mathcal{D}, G: \mathcal{D} \to \mathcal{C}$ are | |
adjoint functors. | |
For any object $c \in \mathcal{C}$, we know that | |
\[ \hom_{\mathcal{D}}(Fc, Fc) \simeq \hom_{\mathcal{C}}(c, GF c), \] | |
so that the identity morphism $Fc \to Fc$ (which is natural in $c$!) corresponds to a map $c \to GFc$ | |
that is natural in $c$, or equivalently a natural | |
transformation | |
\[ \eta: 1_{\mathcal{C}} \to GF. \] | |
Similarly, we get a natural transformation | |
\[ \epsilon: FG \to 1_{\mathcal{D}} \] | |
where the map $FGd \to d$ corresponds to the identity $Gd \to Gd$ under the | |
adjoint correspondence. | |
Here $\eta$ is called the \textbf{unit}, and $\epsilon$ the \textbf{counit.} | |
These natural transformations $\eta, \epsilon$ are not simply arbitrary. | |
We are, in fact, going to show that they determine the isomorphism | |
determine the isomorphism $\hom_{\mathcal{D}}(Fc, d) \simeq | |
\hom_{\mathcal{C}}(c, Gd)$. This will be a little bit of diagram-chasing. | |
We know that the isomorphism $\hom_{\mathcal{D}}(Fc, d) \simeq | |
\hom_{\mathcal{C}}(c, Gd)$ is \emph{natural}. In fact, this is the key point. | |
Let $\phi: Fc \to d$ be any map. | |
Then there is a morphism $(c, Fc) \to (c, d) $ in the product category | |
$\mathcal{C}^{op} \times \mathcal{D}$; by naturality of the adjoint | |
isomorphism, we get a commutative square of sets | |
\[ \xymatrix{ | |
\hom_{\mathcal{D}}(Fc, Fc) \ar[r]^{\mathrm{adj}} \ar[d]^{\phi_*} & \hom_{\mathcal{C}}(c, GF c) | |
\ar[d]^{G(\phi)_*} \\ | |
\hom_{\mathcal{D}}(Fc, d) \ar[r]^{\mathrm{adj}} & \hom_{\mathcal{C}}(c, Gd) | |
}\] | |
Here the mark $\mathrm{adj}$ indicates that the adjoint isomorphism is used. | |
If we start with the identity $1_{Fc}$ and go down and right, we get the map | |
\( c \to Gd \) | |
that corresponds under the adjoint correspondence to $Fc \to d$. However, if we | |
go right and down, we get the natural unit map $\eta(c): c \to GF c$ followed by $G(\phi)$. | |
Thus, we have a \emph{recipe} for constructing a map $c \to Gd$ given $\phi: Fc \to | |
d$: | |
\begin{proposition}[The unit and counit determines everything] | |
Let $(F, G)$ be a pair of adjoint functors with unit and counit transformations | |
$\eta, \epsilon$. | |
Then given $\phi: Fc \to d$, the adjoint map $\psi:c \to Gd$ can be constructed simply as | |
follows. | |
Namely, we start with the unit $\eta(c): c \to GF c$ and take | |
\begin{equation} \label{adj1} \psi = G(\phi) \circ \eta(c): c \to Gd | |
\end{equation} (here $G(\phi): GFc \to Fd$). | |
\end{proposition} | |
In the same way, if we are given $\psi: c \to Gd$ and want to construct a map | |
$\phi: Fc \to d$, we construct | |
\begin{equation} \label{adj2} \epsilon(d) \circ F(\psi): Fc \to FGd \to d. | |
\end{equation} | |
In particular, we have seen that the \emph{unit and counit morphisms determine | |
the adjoint isomorphisms.} | |
Since the adjoint isomorphisms $\hom_{\mathcal{D}}(Fc, d) \to | |
\hom_{\mathcal{C}}(c, Gd)$ and | |
$\hom_{\mathcal{C}}(c, Gd) \to \hom_{\mathcal{D}}(Fc, d) | |
$ | |
are (by definition) inverse to each other, we can determine | |
conditions on the units and counits. | |
For | |
instance, the natural transformation $F \circ \eta$ gives a natural | |
transformation $F \circ \eta: F \to FGF$, while the natural transformation | |
$\epsilon \circ F$ gives a natural transformation $FGF \to F$. | |
(These are slightly different forms of composition!) | |
\begin{lemma} The composite natural transformation $F \to F$ given by | |
$(\epsilon \circ F) \circ (F \circ \eta)$ is the identity. | |
Similarly, the composite natural transformation | |
$G \to GFG \to G$ given by $(G \circ \epsilon) \circ (\eta \circ G)$ is the | |
identity. | |
\end{lemma} | |
\begin{proof} We prove the first assertion; the second is similar. | |
Given $\phi: Fc \to d$, we know that we must get back to $\phi$ applying the | |
two constructions above. The first step (going to a map $\psi: c \to Gd$) is by | |
\eqref{adj1} | |
\( \psi = G(\phi) \circ \eta(c); \) the second step sends $\psi$ to | |
$\epsilon(d) \circ F(\psi)$, by \eqref{adj2}. | |
It follows that | |
\[ \phi = \epsilon(d) \circ F( G(\phi) \circ \eta(c)) = \epsilon(d) \circ | |
F(G(\phi)) \circ F(\eta(c)). \] | |
Now suppose we take $d = Fc$ and $\phi: Fc \to Fc $ to be the identity. | |
We find that $F(G(\phi))$ is the identity $FGFc \to FGFc$, and consequently we | |
find | |
\[ \id_{F(c)} = \epsilon(Fc) \circ F(\eta(c)). \] | |
This proves the claim. | |
\end{proof} | |
\begin{definition} | |
Let $F: \mathcal{C} \to \mathcal{D}, G: \mathcal{D} \to \mathcal{C}$ be | |
covariant functors. An \textbf{adjunction} is the data of two natural | |
transformations | |
\[ \eta: 1 \to GF, \quad \epsilon: FG \to 1, \] | |
called the \textbf{unit} and \textbf{counit}, respectively, such that the | |
composites $(\epsilon \circ F) \circ (F \circ \epsilon): F \to F$ | |
and $(G \circ \epsilon) \circ (\eta \circ G)$ are the identity (that is, the | |
identity natural transformations of $F, G$). | |
\end{definition} | |
We have seen that a pair of adjoint functors gives rise to an adjunction. | |
Conversely, an adjunction between $F, G$ ensures that $F, G$ are adjoint, as | |
one may check: one uses the same formulas \eqref{adj1} and \eqref{adj2} to | |
define the natural isomorphism. | |
For any set $S$, let $F(S)$ be the free group on $S$. | |
So, for instance, the fact that there is a natural map of sets | |
$S \to F(S)$, for any set $S$, and a natural map of | |
groups $F(G) \to G$ for any group $G$, determines the adjunction between the | |
free group functor from $\mathbf{Sets}$ to $\mathbf{Grp}$, and the forgetful | |
functor $\mathbf{Grp} \to \mathbf{Sets}$. | |
As another example, we give a criterion for a functor in an adjunction to be | |
fully faithful. | |
\begin{proposition} \label{adjfullfaithful} | |
Let $F, G$ be a pair of adjoint functors between categories $\mathcal{C}, \mathcal{D}$. | |
Then $G$ is fully faithful if and only if the unit maps $\eta: 1 \to GF$ are | |
isomorphisms. | |
\end{proposition} | |
\begin{proof} | |
We use the recipe \eqref{adj1}. | |
Namely, we have a map $\hom_{\mathcal{D}}(Fc, d) \to | |
\hom_{\mathcal{C}}(c, Gd)$ given by | |
$\phi \mapsto G(\phi) \circ \eta(c)$. This is an isomorphism, since we have an | |
adjunction. | |
As a result, composition with $\eta$ is an isomorphism of hom-sets if and only if $\phi | |
\mapsto G(\phi)$ is an isomorphism. From this the result is easy to deduce. | |
\end{proof} | |
\begin{example} | |
For instance, recall that the inclusion functor from $\mathbf{AbGrp}$ to | |
$\mathbf{Grp}$ is fully faithful (clear). | |
This is a right adjoint to the abelianization functor $G \mapsto G^{ab}$. | |
As a result, we would expect the unit map of the adjunction to be an | |
isomorphism, by \cref{adjfullfaithful}. | |
The unit map sends an abelian group to its abelianization: this is obviously an | |
isomorphism, as abelianizing an abelian group does nothing. | |
\end{example} | |
\subsection{Adjoints and (co)limits} | |
One very pleasant property of functors that are left (resp. right) adjoints is | |
that they preserve all colimits (resp. limits). | |
\begin{proposition} \label{adjlimits} | |
A left adjoint $F: \mathcal{C} \to \mathcal{D}$ preserves colimits. A right | |
adjoint $G: \mathcal{D} \to \mathcal{C}$ preserves limits. | |
\end{proposition} | |
As an example, the free functor from $\mathbf{Sets}$ to $\mathbf{AbGrp}$ is a | |
left adjoint, so it preserves colimits. For instance, it preserves coproducts. | |
This corresponds to the fact that if $A_1, A_2$ are sets, then $\mathbb{Z}[A_1 | |
\sqcup A_2]$ is naturally isomorphic to $\mathbb{Z}[A_1] \oplus | |
\mathbb{Z}[A_2]$. | |
\begin{proof} | |
Indeed, this is mostly formal. | |
Let $F: \mathcal{C}\to \mathcal{D}$ be a left adjoint functor, with right | |
adjoint $G$. | |
Let $f: I \to \mathcal{C}$ be a ``diagram'' where $I$ is a small category. | |
Suppose $\colim_I f$ exists as an object of $\mathcal{C}$. The result states | |
that $\colim_I F \circ f$ exists as an object of $\mathcal{D}$ and can be | |
computed as | |
$F(\colim_I f)$. | |
To see this, we need to show that mapping out of $F(\colim_I f)$ is what we | |
want---that is, mapping out of $F(\colim_I f)$ into some $d \in \mathcal{D}$---amounts to | |
giving compatible $F(f(i)) \to d$ for each $i \in I$. | |
In other words, we need to show that $\hom_{\mathcal{D}}( F(\colim_I f), d) = | |
\lim_I \hom_{\mathcal{D}}( | |
F(f(i)), d)$; this is precisely the defining property of the colimit. | |
But we have | |
\[ \hom_{\mathcal{D}}( F(\colim_I f ), d) = \hom_{\mathcal{C}}(\colim_I f, Gd) | |
= \lim_I \hom_{\mathcal{C}}(fi, Gd) = \lim_I \hom_{\mathcal{D}}(F(fi), d), | |
\] | |
by using adjointness twice. | |
This verifies the claim we wanted. | |
\end{proof} | |
The idea is that one can easily map \emph{out} of the value of a left adjoint | |
functor, just as one can map out of a colimit. | |